Solubilities of l-Proline in Subcritical HFC-134a and Supercritical CO2

Solubility of Mesquite Gum in Supercritical Carbon Dioxide. J. Cruz-Olivares , C. H. Ortiz-Estrada , C. Pérez-Alonso , M. C. Chaparro-Mercado , and C...
0 downloads 0 Views 324KB Size
2780

J. Chem. Eng. Data 2009, 54, 2780–2784

Solubilities of L-Proline in Subcritical HFC-134a and Supercritical CO2 Fluids† Ling Liu, Zhen Li, Zhong-Wen Liu, and Zhao-Tie Liu* Key Laboratory of Applied Surface and Colloid Chemistry (Shaanxi Normal University), Ministry of Education, Xi’an 710062, People’s Republic of China, and School of Chemistry and Materials Science, Shaanxi Normal University, Xi’an 710062, People’s Republic of China

A static method coupled with gravimetric analysis was used to measure the solubility of L-proline in subcritical 1,1,1,2-tetrafluoroethane (HFC-134a) and supercritical CO2 (scCO2). The solubility of L-proline is much higher in HFC-134a fluid than in scCO2. The solubilities of L-proline in HFC-134a and scCO2 are affected by temperature and pressure. The solubilities increase with increasing temperature and pressure for both scCO2 and HFC-134a solvents. The densities of scCO2 and subcritical HFC-134a were calculated using the Peng-Robinson (P-R) equation. Experimentally determined density data were used to assess the accuracy of the P-R equation.

Introduction In recent decades, L-proline, the simplest “protein”,1 has become one of the most attractive molecules in green synthetic chemistry because it is an inexpensive and readily available amino acid. As an effective organocatalyst,2 it has been successfully applied to a variety of reactions3 of which the direct aldol reaction is particularly interesting.4 L-proline suffers from a number of problems that reduce its attractiveness as a catalyst:5 (1) poor solubility in most solvents except water (where it functions as a catalyst, but without enantioselectivity); (2) potential side reactions such as oxazolidinone formation, decarboxylation, and subsequent [3 + 2] cycloaddition reactions; and (3) most of the L-proline-catalyzed aldol reactions have been carried out in organic media such as DMF and DMSO.6 Large amounts of organic solvents are used in chemical reactions, most of which are volatile, toxic, and flammable. From the viewpoint of greener processes, the use of nonhazardous and renewable materials is one of the most important goals of green chemistry. The advantages of supercritical fluids (SCF) compared with conventional liquid solvents include low surface tension, high diffusivity, low viscosity, and high compressibility. In addition, the density, dielectric constant, diffusion coefficient, and solubility parameter can be tuned continuously by varying pressure and temperature. For those reasons supercritical fluids have become attractive solvents for many industrial processes including extraction,7 polymer processing,8 phase transfer reactions and catalysis,9 enzymatic catalysis,10 processing of microelectronic devices,11 and synthesis of nanoparticles.12-15 Supercritical carbon dioxide (scCO2) is the most popular solvent among SCFs due to its nontoxicity, nonflammability, low cost, ready availability, and near-ambient critical temperature. Unfortunately, because CO2 is a nonpolar solvent with weak van der Waals forces it is not suitable for dissolving polar substances. That disadvantage has limited its application to separation, reaction, and material formation processes. Fortunately, polar substances can easily dissolve in fluorohydrocarbons such as chlorodifluoromethane (HCFC-22), trifluo-

Figure 1. Schematic diagram of experimental setup. 1, Carbon dioxide cylinder; 2, ISCO model 260D syringe pump; 3, magnetic stir device; 4, SF-400 high-pressure vessel; 5, pressure transducer; 6, thermocouple assembly; 7, intake valve; 8, back pressure valve; 9, sample vial; 10, reclaimer vase.

romethane (HFC-23), difluoroethane (HFC-32), pentafluoroethane (HFC-125), 1,1,1,2-tetrafluoroethane (HFC-134a), and 1,1,1trifluoroethane (HFC-143a).16-23 Generally, two techniques, the flowing method and the static method, are used to measure solubility in SCFs. Calculations of solubility are correlated using a mathematical model such as the semiempirical model proposed first by Bartle, which was subsequently used successfully by others, the model proposed by Chrastil, and some equations, such as the Peng-Robinson (P-R) and the Soave-Redlich-Kwong (S-R-K) equations.24-29 In the present study, a static method coupled with gravimetric analysis30 was developed for measuring the solubility of solids in scCO2 and HFC-134a. The densities of scCO2 and HFC134a, which are functions of temperature and pressure, were also simulated with the P-R equation. Experimentally determined density data were used to assess the accuracy of the P-R equation, which will provide useful data for future applications of the aldol reaction in SCFs.

Experimental Section * To whom correspondence should be addressed. E-mail: [email protected]. Fax: +86 29 85303682. † Part of the “William A. Wakeham Festschrift”.

Materials and Instruments. CO2 was obtained from Xi’an Yatai Liquid Gas Co., and 1,1,1,2-tetrafluoroethane (HFC-134a)

10.1021/je801009z CCC: $40.75  2009 American Chemical Society Published on Web 04/22/2009

Journal of Chemical & Engineering Data, Vol. 54, No. 9, 2009 2781

P)

RT a V-b V(V + b) + b(V - b)

(2)

The Measurement and Calculation of Pure CO2 and HFC-134a Density. The calculation methods for pure CO2 or HFC-134a density for the measurement system are given in eqs 3 and 4. The density, F, of CO2 or HFC-134a in the high pressure vessel is then

F)

m Vvessel

(3)

F)

n Vvessel

(4)

or

Figure 2. Density of scCO2 calculated by the P-R equation and determined experimentally at 20 MPa and a range of temperatures. 9, experimental data (density of our work); b, P-R equation data (density of P-R equation).

(99.9%) was obtained from Xi’an Jinzhu Modern Chemical Industry Co., Ltd. L-proline (chromatography pure) was purchased from Sinopharm Group Chemical Reagent ShanghaiCo., Ltd. The reagents used in this study were analytically pure grade chemicals. All chemicals were used without further purification. A schematic diagram of the experimental setup for solubility measurements is given in Figure 1. Measurements of Solubility. For each experiment, an excess amount of solute and a small magnetic stirring bar were placed in a glass vial (16 mL) that was then capped with a coarse filter paper attached to the vial with Teflon tape. The sample vial was weighed (W1) and placed inside the pressure vessel. The pressure vessel was sealed and weighed (M1) and then heated to the desired temperature by a heater via a machined internal heating rod. When the desired temperature was reached, stirring was initiated and the vessel was slowly filled with CO2 or HFC-134a until the desired pressure was achieved. The vessel was then reweighed (M2). After sufficient time (usually at least 24 h, as reported previously31) for HFC-134a/solute or CO2/ solute dissolution equilibrium to be attained, the vessel was depressurized and opened. The vial was removed, wiped with a clean tissue, dried, and reweighed (W2). The solubility of the solute was then given by the following equation

Solubility (wt/vol) )

W1 - W2 V1 - V2

(1)

where W1 and W2 are the initial and final mass of solute in the vial, V1 (59 mL) is the volume of the high-pressure vessel, and V2 (16 mL) is the volume of vial. This equation incorporates a correction factor that accounts for precipitation of the solute in the fluid phase in the vial. The volume of the vessel in this equation is the volume accessible to the fluid phase which was determined to be 43 mL. The Foundation of the P-R Equation. Although it has obvious disadvantages, the P-R equation gives an appropriate qualitative description and superior accuracy of quantitative calculation for the phase behavior of supercritical fluids. The P-R equation can apply to many complex systems based on supercritical fluids. In previous work,32 we have derived a P-R equation. The standard form of the P-R equation29,33 is

Results and Discussion Model Analysis. The equations used to calculate standard deviation (s) and absolute error (d) are as follows

s)



∑ (x - jx)2

(5)

n-1 d ) x - jx

(6)

where n - 1 is the degrees of freedom. The density data from the P-R equation and measured for scCO2 at 20 MPa and varying temperature are given in Figure 2. The analysis of these data is shown in Table 1. The standard deviation of the measured density was 0.0515 g · cm-1, and the standard deviation of the density calculated using the P-R equation was 0.0528 g · cm-1. The density data from P-R equation and the data measured for HFC-134a at 19.6 MPa and varying temperature are given in Figure 3. The analysis of these data is shown in Table 2. The standard deviation of the measured density was 0.0474 g · cm-1, and the standard deviation of the density calculated by P-R equation was 0.0564 g · cm-1. At the same pressure, the density curve from the P-R equation for scCO2 or HFC-134a is nearly parallel to the corresponding measured density curve. The two standard deviations are very similar which indicates that the P-R equation is accurate for use in our experimental calculation. The density from the P-R equation and the density measured for scCO2 at 313 K and varying pressure are shown in Figure 4. The analysis of the density data is shown in Table 3. The curve of measured density tends toward the curve of density calculated by the P-R equation when the pressure is between 15 and 25 MPa. The two curves deviate from each other gradually at lower and higher pressures. The standard deviation Table 1. Standard Deviation and Absolute Error of CO2 Density Calculated Using the P-R Equation and Experimentally Determined at 20 MPa and a Range of Temperaturesa T

F1

F2

d1

d2

s1

s2

K

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

308 313 318 323 328 333

0.87 ( 0.06 0.85 ( 0.04 0.82 ( 0.01 0.79 ( 0.01 0.76 ( 0.04 0.73 ( 0.07

0.8669 0.8412 0.8143 0.7860 0.7565 0.7258

0.0654 0.0434 0.0147 -0.0116 -0.0403 -0.0715

0.0684 0.0427 0.0158 -0.0124 -0.0419 -0.0726

0.0515

0.0528

a F1, experimentally determined density; F2, density from P-R equation; d1, absolute error of experimental density; d2, absolute error of density from P-R equation; s1, standard deviation of experimental density; s2, standard deviation of P-R equation density.

2782

Journal of Chemical & Engineering Data, Vol. 54, No. 9, 2009 Table 3. Standard Deviation and Absolute Error of scCO2 Density Calculated Using the P-R Equation and Experimentally Determined at 313 K and a Range of Pressuresa P

F1

F2

d1

d2

s1

s2

MPa

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

10 15 20 25 36

0.57 ( 0.22 0.78 ( 0.01 0.83 ( 0.03 0.89 ( 0.09 0.90 ( 0.10

0.6373 0.7824 0.8412 0.8802 0.9400

-0.2189 -0.0137 0.0349 0.0949 0.1029

-0.1789 -0.0338 0.0250 0.0640 0.1238

0.1313

0.1153

a F1, experimentally determined density; F2, density from P-R equation; d1, absolute error of experimental density; d2, absolute error of density from P-R equation; s1, standard deviation of experimental density; s2, standard deviation of P-R equation density.

Figure 3. Density of HFC-134a calculated using the P-R equation and determined experimentally at 20 MPa and a range of temperatures. 9, experimental data (density of our work); b, P-R equation data (density of P-R equation). Table 2. Standard Deviation and Absolute Error of HFC-134a Density Calculated Using the P-R Equation and Experimentally Determined at 19.6 MPa and a Range of Temperaturesa T

F1

F2

d1

d2

s1

s2

K

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

288 298 308 318 328

1.29 ( 0.06 1.27 ( 0.03 1.24 ( 0.002 1.21 ( 0.03 1.18 ( 0.06

1.3146 1.2879 1.2600 1.2309 1.2004

0.0592 0.0306 0.0018 -0.0323 -0.0592

0.0655 0.0388 0.0109 -0.0667 -0.0487

0.0474

0.0564

a F1, experimentally determined density; F2, density from P-R equation; d1, absolute error of experimental density; d2, absolute error of density from P-R equation; s1, standard deviation of experimental density; s2, standard deviation of P-R equation density.

Figure 5. Density of HFC-134a calculated using the P-R equation and determined experimentally at 308 K and a range of pressures. 9, experimental data; 2, P-R equation data. Table 4. Standard Deviation and Absolute Error of HFC-134a Density Calculated Using the P-R Equation and Experimentally Determined at 308 K and a Range of Pressuresa P

F1

F2

d1

d2

s1

s2

MPa

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

g · cm-3

5.0 10.1 15.4 19.6 23.7

1.15 ( 0.05 1.18 ( 0.03 1.21 ( 0.002 1.24 ( 0.03 1.26 ( 0.05

1.1712 1.2088 1.2383 1.2600 1.2782

-0.0536 -0.0299 0.0023 0.0291 0.0523

-0.0601 -0.0225 0.0070 0.0287 0.0469

0.0429

0.0424

a F1, experimentally determined density; F2, density from P-R equation; d1, absolute error of experimental density; d2, absolute error of density from P-R equation; s1, standard deviation of experimental density; s2, standard deviation of P-R equation density.

Figure 4. Density of scCO2 calculated using the P-R equation and determined experimentally at 313 K and a range of pressures. 9, experimental data (density of our work); 2, P-R equation data (density of P-R equation).

of the measured density was 0.1313 g · cm-1, and the standard deviation of density calculated by P-R equation was 0.1153 g · cm-1. The density from the P-R equation and the measured density for HFC-134a at 308 K at varying pressures are shown in Figure 5. The analysis of the density data is shown in Table 4. The standard deviations of the measured density and the density calculated from the P-R equation were 0.0429 and 0.0424 g · cm-1, respectively. At the same temperature, the density curve from the P-R equation for HFC-134a is nearly parallel to the measured density curve. The two standard

deviations are very similar, which indicates that the P-R equation is sufficiently accurate for our experimental calculation. There are systematic errors between data calculated by the P-R equation and data obtained in our experimental calculation. We hypothesize that the systematic errors are from estimated errors on the pressure and temperature settings. The temperature was controlled during each experiment with a variation of (0.5 K). The pressure was measured using a pressure transducer (Beijing, Zhengkai, MCYB) and controlled during each experiment with a variation of (0.01 MPa). The other errors are from the experimentally determined weights. Solubility Measurement. A static method coupled with gravimetric analysis was developed for measuring the solubility of L-proline in scCO2 and HFC-134a. Figure 6 reveals that the solubility of L-proline in HFC-134a increases with increase in temperature at constant pressure (19.6 MPa). The solubility of L-proline in scCO2 at 20 MPa and

Journal of Chemical & Engineering Data, Vol. 54, No. 9, 2009 2783

Figure 6. Solubility of L-proline in HFC-134a at 19.6 MPa and a range of temperatures. Error bars use the 95 % confidence interval, which corresponds to approximately 3 standard deviations taken from triplicate runs. 9, solubility data.

Figure 7. Solubility of L-proline in scCO2 at 20 MPa and a range of temperatures. Error bars use the 95 % confidence interval, which corresponds to approximately 3 standard deviations taken from triplicate runs. 9, solubility data.

varying temperature is shown in Figure 7. The error bars correspond to one standard deviation calculated from three replicate measurements done at each temperature and pressure. The solubility increased with increasing temperature at constant pressure for L-proline in HFC-134a and scCO2, because when the temperature increases, compared with the decrease of density, the increase of vapor pressure is the most important factor. This explanation applies to both CO2 and HFC-134a solvents. The solubility of L-proline in HFC-134a at a fixed temperature (308 K) and varying pressure (5 MPa to 25 MPa) is given in Figure 8. The solubility of L-proline in scCO2 at a fixed temperature (313 K) and varying pressure (10 MPa to 36 MPa) is given in Figure 9. The error bars correspond to one standard deviation calculated from triplicate measurements at each temperature and pressure. The solubility of a solute in a fluid depends strongly on the density and increases with increase in density hence with increase in pressure. This is in accord with conventional wisdom that the density of a supercritical fluid must increase in order to increase the solubility and extraction efficiency.34 Considering the thermodynamic view, at a given temperature the Hildebrand formula ∆HM ) V(δ1-δ2) predicts

Figure 8. Solubility of L-proline in HFC-134a at 308 K and a range of pressures. Error bars use the 95 % confidence interval, which corresponds to approximately 3 standard deviations taken from triplicate runs. 9, solubility data.

Figure 9. Solubility of L-proline in scCO2 at 313 K and a range of pressures. Error bars use the 95 % confidence interval, which corresponds to approximately 3 standard deviations taken from triplicate runs. 9, solubility data. Table 5. Experimental Solubility Data for L-Proline in scCO2 and HFC-134a at 20 MPa and 308 K, 318 K, and 328 Ka CO2 P/MPa T/K 20 a

m/g

308 0.0120 318 0.0327 328 0.0462

HFC-134a m/g

CO2

HFC-134a -1

S × 10 /g · mL

0.2232 0.3211 0.4134

4

2.79 7.60 10.74

S × 104 /g · mL-1 51.90 74.70 96.20

m, amount dissolved; S, solubility

that if the solubility parameters of a fluid solvent δ1 and a solute δ2 are almost equal in magnitude, the enthalpy of mixing, ∆HM, is small and the solute should dissolve in the solvent. The results indicate that at low pressure (10 MPa to 15 MPa) the increasing trend of solubility with increase in pressure is slow, while at higher pressure (15 MPa to 20 MPa) the increasing trend is fast. Pressure has the predominant influence on the density of a supercritical fluid; the density increases with increasing pressure at constant temperature. The solubilities of L-proline in scCO2 and HFC-134a at the same pressure are listed in Table 5. Under the same conditions, the solubility of L-proline in HFC-134a is much higher than in scCO2. The results indicate that HFC-134a is a more polar

2784

Journal of Chemical & Engineering Data, Vol. 54, No. 9, 2009

solvent than scCO2 for dissolving the polar compound. Because scCO2 is a nonpolar solvent with weak van der Waals forces, between solute and scCO2 there are dipole-quadrupole forces that inhibit dissolution of the solute.35 Consequently, polar solutes cannot easily dissolve in scCO2. However, HFC-134a is a polar solvent that has a strong solvent power for polar solutes. Since the dipole-dipole force between polar solute and HFC-134a promotes dissolution, the solubility of L-proline in HFC-134a is higher than in scCO2.32

Conclusion A static method coupled with gravimetric analysis was developed for measuring the solubility of L-proline in scCO2 and HFC-134a fluids. The solubility increases with increasing temperature and pressure for scCO2 and HFC-134a. The solubility of the polar solute in HFC-134a was found to be much higher than in the most commonly used supercritical solvent, that is, CO2. The densities of CO2 and HFC-134a calculated using the P-R equation are close to the experimental values in a wide range of temperatures and pressures.

Literature Cited (1) List, B. The ying and yang of asymmetric aminocatalysis. Chem. Commun. 2006, 819–824. (2) (a) Kocˇovsky´, P.; Malkov, A. V. Organocatalysis in organic synthesis. Tetrahedron 2006, 62, 255–255. (b) Marigo, M.; Jørgensen, K. A. Organocatalytic direct asymmetric a-heteroatom functionalization of aldehydes and ketones. Chem.Commun. 2006, 2001–2011. (3) (a) Sabitha, G.; Fatima, N.; Reddy, E. V.; Yadav, J. S. First examples of proline-catalyzed domino Knoevenagel/Hetero-Diels-Alder/Elimination reactions advanced synthesis & catalysis. AdV. Synth. Catal. 2005, 347, 1353–1355. (b) Chen, S. H.; Hong, B. C.; Su, C. F.; Sarshar, S. An unexpected inversion of enantioselectivity in the proline catalyzed intramolecular Baylis-Hillman reaction. Tetrahedron Lett. 2005, 46 (51), 8899–8903. (c) Chowdari, N. S.; Ramachary, D. B.; Barbas, C. F. III. Organocatalysis in ionic liquids: highly efficient L-prolinecatalyzed direct asymmetric mannich reactions involving ketone and aldehyde nucleophiles. Synlett. 2003, 1906–1909. (d) Notz, W.; Tanaka, F.; Watanabe, S.; Chowdari, N. S.; Turner, J. M.; Thayumanavan, R.; Barbas, C. F., III. The direct organocatalytic asymmetric mannich reaction: unmodified aldehydes as nucleophiles. J. Org. Chem. 2003, 68, 9624–9634. (4) (a) Chandrasekhar, S.; Reddy, N. R.; Sultana, S. S.; Narsihmulu, Ch.; Reddy, K. V. L-Proline catalysed asymmetric aldol reactions in PEG400 as recyclable medium and transfer aldol reactions. Tetrahedron 2006, 62, 338–345. (b) Bernard, A. M.; Frongia, A.; Guillot, R.; Piras, P. P.; Secci, F.; Spiga, M. L-Proline-catalyzed direct intermolecular asymmetric aldol reactions of 1-phenylthiocycloalkyl carboxaldehydes with ketones. Easy access to spiro- and fused-cyclobutyl tetrahydrofurans and cyclopentanones. Org. Lett. 2007, 9 (3), 541–544. (5) Pihko, P. M.; Laurikainen, K. M.; Usano, A.; Nyberg, A. I.; Kaavi, J. A. Effect of additives on the proline-catalyzed ketone-aldehyde aldol reactions. Tetrahedron 2006, 62, 317–328. (6) Bellis, E.; Kokotos, G. Proline-modified poly(propyleneimine) dendrimers as catalysts for asymmetric aldol reactions. J. Mol. Catal. A: Chem. 2005, 241, 166–174. (7) Eckert, C. A.; Knutson, B. L.; Debenedetti, P. G. Supercritical fluids as solvents for chemical and materials processing. Nature 1996, 383, 313–318. (8) Cooper, A. I. Polymer synthesis and processing using supercritical carbon dioxide. J. Mater. Chem. 2000, 10, 207–234. (9) Dillow, A. K.; Yun, S. L.; Suleiman, D. S.; Boatright, D. L.; Liotta, C. L.; Eckert, C. A. Kinetics of a phase-transfer catalysis reaction in supercritical fluid carbon dioxide. Ind. Eng. Chem. Res. 1996, 35, 1801–1806. (10) Holmes, J. D.; Steytler, D. C.; Rees, G. D.; Robinson, B. H. Bioconversions in a water-in-CO2 microemulsion. Langmuir 1998, 14, 6371–6376. (11) Sundararajan, N.; Yang, S.; Ogino, K.; Valiyaveettil, S.; Wang, J.; Zhou, X.; Ober, C. K.; Obendorf, S. K.; Allen, R. D. Supercritical CO2 processing for submicron imaging of fluoropolymers. Chem. Mater. 2000, 12, 41–48. (12) Ji, M.; Chen, X.; Wai, C. M.; Fulton, J. L. Synthesizing and dispersing silver nanoparticles in a water-in-supercritical carbon dioxide microemulsion. J. Am. Chem. Soc. 1999, 121, 2631–2632.

(13) Ohde, H.; Ohde, M.; Bailey, F.; Kim, H.; Wai, C. M. Water-in-CO2 microemulsions as nanoreactors for synthesizing CdS and ZnS nanoparticles in supercritical CO2. Nano Lett. 2002, 2, 721–724. (14) Ohde, H.; Rodriguez, J. M.; Ye, X. R.; Wai, C. M. Synthesizing silver halide nanoparticles in supercritical carbon dioxide utilizing a water in-CO2 microemulsion. Chem. Commun. 2000, 23, 2353–2354. (15) Dong, X.; Potter, D.; Erkey, C. Synthesis of CuS nanoparticles in water-in-carbon dioxide microemulsions. Ind. Eng. Chem. Res. 2002, 41, 4489–4493. (16) Lim, J. S.; Park, J. Y.; Yoon, C. H.; Lee, Y. W.; Yoo, K. P. Cloud points of poly(L-lactide) in HCFC-22, HFC-23, HFC-32, HFC-125, HFC-143a, HFC-152a, HFC-227ea, dimethyl ether (DME), and HCFC22 + CO2 in the supercritical state. J. Chem. Eng. Data 2004, 49, 1622–1627. (17) Eastoe, J.; Gold, S. Self-assembly in green solvents. Phys. Chem. Chem. Phys. 2005, 7, 1352–1362. (18) Olsen, S. A.; Tallman, D. E. Conductivity and voltammetry in liquid and supercritical halogenated solvents. Anal. Chem. 1996, 68, 2054– 2061. (19) Abbott, A. P.; Eardley, C. A.; Harper, J. C.; Hop, E. G. Electrochemical investigations in liquid and supercritical 1,1,1,2-tetrafluoroethane (HFC-134a) and difluoromethane (HFC-32). J. Electroanal. Chem. 1998, 457, 1–4. (20) Tackson, K.; Fulton, J. L. Microemulsions in supercritical hydrochlorofluorocarbons. Langmuir 1996, 12, 5289–5295. (21) Butz, N.; Porte, C.; Courrier, H.; Krafft, M. P.; Vandamme, Th. F. Reverse water-in-fluorocarbon emulsions for use in pressurized metered-dose inhalers containing hydrofluoroalkane propellants. Int. J. Pharm. 2002, 238, 257–269. (22) Eastoe, J.; Thorpe, A. M.; Eastone, J.; Dupont, A.; Heenan, R. K. Microemulsion formation in 1,1,1,2-tetrafluoroethane (134a). Langmuir 2003, 19, 8715–8720. (23) Bush, D.; Echert, C. A. Prediction of solid-fluid equilibria in supercritical carbon dioxide using linear solvation energy relationships. Fluid Phase Equilib. 1998, 150-151, 479–492. (24) Ozcan, A. S.; Clifford, A. A.; Bartle, K. D. Solubility of disperse dyes in supercritical carbon dioxide. J. Chem. Eng. Data 1997, 42, 590– 592. (25) Joung, S. N.; Yoo, K. Solubility of disperse anthraquinone and azodyes in supercritical carbon dioxide at 313.15 to 393.15 K and from 10 to 25 MPa. J. Chem. Eng. Data 1998, 43, 9–12. (26) Lee, J. W.; Min, J. M.; Bae, H. K. Solubility measurement of disperse dyes in supercritical carbon dioxide. J. Chem. Eng. Data 1999, 44, 684–687. (27) Yamini, Y.; Bahramifar, N. Solubility of polycyclic aromatic hydrocarbons in supercritical carbon dioxide. J. Chem. Eng. Data 2000, 45, 53–56. (28) Chrastil, J. Solubility of solids and liquids in supercritical gases. J. Phys. Chem. 1982, 86, 3016–3021. (29) Peng, D.; Robinson, D. B. A new two-constant equation of state. Ind. Eng. Chem. Fundam. 1976, 15, 59–64. (30) Sherman, G.; Shenoy, S.; Weiss, R. A.; Erkey, C. A Static Method Coupled with Gravimetric Analysis for the Determination of Solubilities of Solids in Supercritical Carbon Dioxide. Ind. Eng. Chem. Res. 2000, 39, 846–848. (31) Liu, Z. T.; Liu, L.; Wu, J.; Song, L. P.; Gao, Z.; Dong, W.; Lu, J. Solubility and phase behaviors of AOT analogue surfactants in 1,1,1,2tetrafluoroethane and supercritical carbon dioxide. J. Chem. Eng. Data 2006, 51, 2045–2050. (32) Liu, Z. T.; Wu, J.; Liu, L.; Sun, C.; Song, L. P.; Gao, Z.; Dong, W.; Lu, J. Solubilities of AOT analogues surfactants in supercritical CO2 and HFC-134a fluids. J. Chem. Eng. Data 2006, 51, 1761–1768. (33) Valderrama, J. O.; Alvarez, V. H. Phase equilibrium in supercritical CO2 mixtures using a modified Kwak-Mansoori mixing rule. AIChE J. 2004, 50, 480–488. (34) Miller, D. J.; Hawthorne, S. B. Determination of solubilities of organic solutes in supercritical CO2 by on-Line flame ionization detection. Anal. Chem. 1995, 67, 273–279. (35) Abbott, A. P.; Corr, S.; Durling, N. E.; Hope, E. G. Solubility of substituted aromatic hydrocarbons in supercritical difluoromethane. J. Chem. Eng. Data 2002, 47, 900–905. Received for review December 31, 2008. Accepted March 30, 2009. The authors gratefully acknowledge the financial support of the Specialized Research Fund for the Doctoral Program of Higher Education (20070718003) and the Natural Science Foundation of Shaanxi Province (2007B07).

JE801009Z